throbber
. .
`
`.- .
`
`Molten Glass Corrosion Resistance of Immersed Combustion-Heating
`Tube Materials in E-Glass
`
`S. Kamakshi Sundaram,. Jen-Yen Hsu,' and Robert E Speyer.
`School of Materials Science and Engineering, Georgia Institute of Technology, Atlanta, Georgia
`30332-0245
`
`J Am. Cerum. SOC., 78 171 1940-46 (1995)
`
`The corrosion resistance of molybdenum, molybdenum
`disilicide, and a SiC(,,/AI,O, composite to molten E-glass at
`1550°C was studied. Mo showed no tendency to oxidize as it
`was immersed in soda-lime silicate glass in a parallel study.
`MoSi, was corroded by soluble molecular oxygen, leaving a
`Mo,Si, interface behind. The SiC(,,/AI,O, composite was
`corroded at a more rapid rate wherein the S i c component
`was oxidized to form amorphous silica and CO bubbles.
`Based on these results, the activity of soluble molecular
`oxygen in E-glass was determined to be in the range of 2.4 X
`10-14 to 2.0 x 10-8.
`
`I. Introduction
`HE work discussed herein is part of a larger research thrust
`
`T to develop materials for immersed gas-fired radiant burner
`
`tubes for glass melters. In the evaluation of candidate materials,
`our research has adopted a parallel approach, separately evalu-
`ating molten glass corrosion and combustion gas corrosion, as
`well as using thermodynamic simulations for both processes.
`Gas corrosion investigations are reported elsewhere.',' In the
`present paper, the corrosion resistance of candidate materials to
`a molten E-glass (fiberglass) composition is presented.
`Of the candidate materials evaluated, those presented here
`are molybdenum, molybdenum disilicide, and a SiC,,,/Al,O,
`composite ( p refers to particulate). Molybdenum has been used
`as the electrode material for electric melting and boosting in
`glass-melting tanks for many year^.^,^ Molybdenum disilicide
`has demonstrated a self-protective mechanism of forming an
`amorphous SiO, layer at the surface in an oxidizing atmo-
`sphere.' In the case of the SiC,,,/AI,O, composite, the combina-
`tion of high thermal conductivity of Sic and the refractory
`properties of A1,0, was anticipated to make it suitable for the
`combustion product side of the present application. This work
`closely parallels corrosion studies of these candidate materials
`in soda-lime-silicate glass at 1565°C.3 In that work, all three
`materials were found to corrode by oxidation reactions. An
`oxidation product interfacial region of MOO,(,, for Mo, and
`Mo,Si, for MoSi,, formed, which periodically cracked away
`due to interfacial stresses. CO/CO, bubbles formed as a reaction
`product in the corrosion of the SiC(,,/Al,O, composite.
`
`from MoSi, or SiC~,,/A1,03, alloys with platinum to destroy the
`crucible. Therefore, fusion-cast AZS (UNICORSOI , Corhart
`Refractories, Louisville, KY) crucibles with containment vol-
`umes of 3.81 cm diameter by 3.81 cm depth were used. AZS
`consists of 50.80 wt% A1,0,, 32.60 wt% ZrO,, 14.90 wt%
`SO,, 1.25 wt% Na,O, 0.13 wt% Fe,O,, and 0.12 wt% Ti0,.9
`Molybdenum (Johnson Matthey/AESAR, Ward Hill, MA)
`specimens were of 12.5 mm diameter by 12.5 mm length for
`the purpose of complete immersion. Chemical analysis for trace
`elements in as-received molybdenum showed less than 1 ppm
`of Al, Ca, Cr, Cu, Mg, Mn, Ni, Pb, Si, Sn, and Ti, less than
`2 ppm of C and 0, and less than 14 ppm of Fe. Cylindrical
`specimens of MoSi, (Kanthal Super 33, Kanthal, Bethel, CT)
`and SiC~,,/AI,O, (DuPont-LanxideTM Composites, Wilmington,
`DE) of 12.5 mm diameter by 60 mm length were used. As
`determined by scanning electron microscopy (SEM) image
`analysis, the as-received MoSi, contained 1.7 vol% Mo,Si,
`grains, generally in contact with amorphous aluminosilicate
`glassy pockets, which in turn comprised 18.6 vol% of the
`specimen.
`The E-glass (Owens/Corning Fiberglas, Granville, OH) used
`in the present investigation is a common fiberglass composition.
`A chemical analysis of the as-received glass, in weight percent,
`shows SiO, = 54.4%, CaO = 17.9%, A1,0, = 14.7%, B,O, =
`6.3%, MgO = 4.7%, Na,O = 0.6%, TiO, = 0.6%, F, = 0.5%,
`and Fe,O, = 0.33%.
`The as-received glass was crushed to -40 mesh, and the
`AZS crucible with 48.5 g of glass was placed inside an electri-
`cally heated furnace. The glass was heated to 1550°C and
`allowed to equilibrate at that temperature for 2 h to obtain
`bubble-free glass. Test specimens were then loaded into the
`crucible and exposed for 12, 24, or 48 h. Details of measure-
`ment and characterization procedures may be found in related
`published
`
`11. Experimental Procedure
`The ASTM standard method C621-84 for isothermal corro-
`sion resistance of refractories to molten glass was revised for
`the present in~estigation.~ Since molybdenum oxidizes and vol-
`atilizes above the glass line at temperatures exceeding -6OO"C,
`it was fully immersed. Soluble silicon in the melt, originating
`
`J. L. Smialek-contributing editor
`
`"L 0
`
`10-
`
`20
`30
`Time (hrs)
`
`40
`
`50
`
`Manuscript No. 194582. Received May 10,1993; approved February 16, 1995.
`Supported by the Gas Research Institute under Contract No. 5090-298-2073. Much
`of this work was performed at the New York State College of Ceramics at Alfred
`University.
`'Member, American Ceramic Society.
`
`Fig. 1. Corrosion of refractory containment materials at glass-line
`(solid lines) and half-down (dotted lines) regions. The error bars indi-
`cate the standard deviation in the measurement of specimen diameters
`before the corrosion tests.
`
`1940
`
`GE-1022.001
`
`

`

`July 1995
`
`Molten Glass Corrosion Resistance of Immersed Combustion-Heating Tube Materials in E-Glass
`
`1941
`
`Table I. Chemical Analyses (wt%) of Glass after Interaction with AZS at 1550°C
`Corrosion time (h) = 0
`24
`3
`6
`12
`36
`48
`54.4
`53.8
`54.5
`54.8
`53.5
`54
`53.2
`15.7
`15.6
`15.7
`17.4
`17.2
`18.1
`14.7
`0.15
`0.55
`1.01
`0
`0
`0
`0
`
`SiO,
`A1203
`ZrO,
`
`111. Results and Discussion
`The choice of A Z S was made after static corrosion testing
`of fusion-cast A Z S , MgO-partially-stabilized zirconia (PSZ)
`(Coors Ceramics, Golden, CO), fusion-cast chromia (Carborun-
`dum, Monofrax Refractories Division, Falconer, NY), and
`bonded chromia (Corhart Refractories, Buchannon, WV)
`refractories, following the ASTM C621-84 standard. The
`results of corrosion testing, presented in Fig. 1, indicate that
`AZS had an inferior corrosion resistance, both at the glass line
`and half-down regions, to bonded and fusion-cast chromia. The
`half-down corrosion rates of bonded chromia and PSZ could
`not be measured due to specimen cracking, for the latter,
`undoubtedly from destabilization of the cubic crystal structure.
`Fusion-cast chromia demonstrated slight recession up to 24 h,
`followed by an apparent increase in cross-sectional area after
`48 h of exposure. This remarkable feature consistently appeared
`during three separate trials, in both half-down and glass-line
`measurements. Two possibilities are suggested: (1) The high
`vapor pressure of chromia tended to expand porous regions and
`increased specimen dimensions by creep. (2) The chemistry of
`solid solutions within the refractory changed via incorporation
`of ions from the glass into the chromia structure, in turn causing
`a volume expansion. Bonded chromia did not show this dilation
`behavior. AZS refractory was, nevertheless, selected as the
`containment material over chromia refractories, to eliminate
`glass coloration.
`The chemical analyses of the glass after interaction with AZS
`for varying times are shown in Table I. No significant changes
`in these constituents were observed up to 12 h of exposure.
`Beyond 12 h, the A1,0, and ZrO, content increased; after 48 h
`of exposure, the A1,0, weight fraction increased to 2.4%, and
`ZrO, to 1.01%. The effect of this compositional change on the
`specimen corrosion rate is not known. Replenishing glass melt
`after every 12 h of exposure was considered and rejected, since
`it changes the present static testing to a semidynamic corrosion
`test.
`Figures 2(a and b) show the corrosion of cylindrical speci-
`mens by molten glass (in turn contained in AZS crucibles). In
`the case of glass-line corrosion, SiC,,,/Al,O, showed a corro-
`sion rate comparable to that of MoSi, for the first 24 h but
`accelerated beyond that. SiC,,J,/A1,O, corrosion appeared
`
`largely independent of positioning at the glass-line or half-
`down locations. MoSi, showed an order of magnitude decrease
`in recession rate at the half-down position as compared to the
`glass-line position. Immersed Mo showed the least corrosion,
`with no measurable corrosion for the first 24 h.
`(1) Molybdenum
`Figure 3 shows the Mwglass interfacial morphology; no
`interfacial layer can be seen. This is corroborated by a line EDS
`scan (not shown) in which a sharp decrease in Mo concentration
`with corresponding sharp increase in Si concentration was
`observed at the interface. XRD patterns starting from the glass
`side, penetrating into the bulk (not shown), did not show any
`interfacial phase at the Mo-glass interface.
`Neither MOO, nor MOO,,,, were detected at the interfacial
`region. For the equilibria,
`
`AGY,,,, = -268.6 kJ/mol, ao,,,,,,,, = 2.0 X
`Thus, the
`molecular oxygen activity in the molten glass would have to be
`greater than or equal to 2.0 X lo-' for this reaction to be
`favorable. The oxygen activity would have to equal or exceed
`7.4 X
`for MOO,,,, to be favorable.
`The concentration of molecular oxygen in alkali silicate
`melts has been shown" to be 1% of that in the atmosphere,
`which translates to a soluble molecular oxygen activity of -2 X
`lo-, for glass in an air atmosphere. This was consistent with an
`observed MOO, surface layer on Mo exposed to molten soda-
`lime-silicate glass,, since this oxygen activity was well above
`the critical value of 2.0 X lo-*. No information is available for
`the solubility of molecular oxygen in calcium aluminosilicates
`such as the E-glass composition studied herein. However, based
`on the fact that no molydenum oxides were observed, an oxy-
`gen activity below 2.0 X lo-' at 1550°C in E-glass is indicated,
`assuming that there were no kinetic hindrances to oxide
`formation.
`The enhancement in Mo corrosion between 24 and 48 h of
`exposure is suspected to be a result of solution of impurity
`phases from the AZS crucible. The presence of impurity phases
`such as As,O,, Na,SO, have been shown" to enhance the corro-
`sion of molybdenum in molten glass, but these phases are not
`present in AZS. However, solution of zirconia, iron oxide and
`
`C .z fa
`2 - 1
`
`.......................................
`
`.......' A t
`
`
`
`10
`
`40
`
`50
`
`"- 0
`
`10
`
`20
`30
`30
`20
`Time (hrs)
`Time (hrs)
`(a)
`(b)
`Fig. 2.
`(a) Specimen corrosion at the glass-line (dotted lines) and half-down regions (solid lines). (b) Magnified view of corrosion of half-down Mo
`and MoSi,. Propagated error calculations' result in error bars smaller than the symbol size and are therefore not shown. The dashed line shows radius
`reduction results corrected for the thickness of the original vitreous silica layer on as-received MoSi, (see text for discussion).
`
`40
`
`so
`
`GE-1022.002
`
`

`

`1942
`
`Journal of the American Ceramic Society-Sundaram et al.
`
`Vol. 78, No. 7
`
`Fig. 3. Molybdenum-glass interfacial region after 48 h of corrosion. M: molybdenum, G: glass. Left: secondary electron image. Right: backscat-
`tered image.
`
`titania, present in AZS, would be expected to alter the solubility
`of molecular oxygen in the glass. It is therefore possible that
`corrosion of the AZS crucible would cause the corrosion of Mo
`by increasing the solubility of molecular oxygen in E-glass.
`(2) Molybdenum Disilicide
`The as-received MoSi, had a fused silica surface layer of
`-10 pm thickness. After preheat treatment at 1000°C for
`10 min, the surface layer became somewhat thinner and discon-
`tinuous. If it is assumed that this vitreous coating mixed with
`the glass so that it was effectively removed, then the measured
`postcorrosion specimen radii would be affected, but not the
`trend in radius recession with exposure time. The dashed line
`in Fig. 2(b) shows the measured MoSi, specimen radii after
`subtraction of a 10 pm surface layer.
`Figure 4 shows XRD patterns starting from the glass, pene-
`trating through the interface, into the bulk. From the nearby
`glass to the interfacial region, molybdenum was identified. The
`major interfacial phase was identified to be Mo,Si,. Minor
`
`quantities of Mo,Si could not be confirmed or denied, since the
`most intense peaks of this phase overlap with those of Mo,Si,.
`An SEM EDS pattern was taken (not shown) of the surface
`from which the XRD second from the front in Fig. 4 was taken.
`Compared to as-received MoSi,, appreciable molybdenum
`enhancement relative to silicon was apparent, corresponding to
`an XRD pattern showing predominantly Mo,Si,.
`Figures 5(a and b) show the MoSi2-glass interfacial features
`at the glass-line and half-down regions, respectively. The
`interfacial layer had a thickness of 133.2 ? 4.3 pm at the half-
`down region, and 13.0 ? 3.2 pm at the glass-line region. From
`the backscattered images, the interfacial region was richer in
`atomically heavier atoms. Since oxygen is lighter than both
`metals, the interfacial region is not an oxide such as MOO, or
`SiO,. Mo being heavier than SI, the interfacial region is inter-
`preted to be Mo-rich. Further corroboration is provided by
`EDS line scans (Fig. 6). At the interface, the Si-to-Mo ratio
`is decreased. Although the Si intensity varies throughout the
`interfacial region, a general Si concentration gradient from
`
`8 0 6
`Fig. 4. XRD patterns as a function of penetration, analyzing glass, interfacial region, and molybdenum disilicide bulk. m: molybdenum, 1: MoSi,,
`and 2: Mo,Si,. XRD surfaces were prepared by repeatedly grinding a flat glass-exposed surface with 600 pm abrasive paper for 10 min and cleaning.
`
`GE-1022.003
`
`

`

`July 1995
`
`Molten Glass Corrosion Resistance of Immersed Conzbustion-Heating Tube Materials in E-Glass
`
`1943
`
`Molybdenum disilicide-glass interfacial region at (a) glass-line and (b) half-down regions after 48 h of corrosion. M: molybdenum
`Fig. 5.
`disilicide, I: interface, F: flaked-off layer, and G: glass. Left: secondary electron image. Right: backscattered image.
`
`interior to exterior is apparent. The fluctuations in Si and Mo
`concentrations in the MoSi, bulk, as well as the interfacial
`region, are attributed to dispersed pockets of amorphous alumi-
`nosilicate phase throughout the material. A small Mo EDS peak
`away from the interface suggests a limited molybdenum-rich
`waste stream into the glass. As implied from Fig. 5, at the glass
`line, the flaked-off reaction-product layer was drawn away from
`the interface by enhanced glass-line convection currents.” This
`corrosion product layer remained close to the interface at the
`half-down region. This also explains the decreased interfacial
`layer thickness at the glass line as compared to the half-down
`region.
`XRD, SEM, and EDS results provide conclusive evidence
`that the mechanism of MoSi, corrosion is removal of silicon
`from MoSi, to the glass, leaving a silicon-deficient molybde-
`num silicide interface. This is an oxidation reaction where oxy-
`gen was provided from soluble molecular oxygen: $MoSi, +
`= ?Mo,Si, + SiO,. Figure 7 is an Ellingham diagram
`02(g,ars)
`
`for equilibria of interest among compounds of Mo, Si, and 0.
`As implied by the lowest standard free-energy equilibria line,
`formation of silica from MoSi, via the most oxygen conserva-
`tive mechanism is favored-silica
`is formed by oxidation and
`removal of silicon from MoSi, to form Mo,Si,, rather than
`complete oxidation of the silicon component, the molybdenum
`component, or both. In addition, the most favored reaction is
`that most conservative in silicon removal from the compound;
`e.g., Mo,Si, formed rather than Mo,Si or Mo. If no MoSi, is
`locally available, then it would be thermodynamically favorable
`for Mo,Si, to convert to Mo,Si, and by the same argument, for
`Mo,Si to convert to Mo.
`These predictions are consistent with the above experimental
`results where, at the interface, Mo,Si, formed. This interface
`would not be expected to convert to Mo,Si and Mo if MoSi, is
`locally available, since the latter has a greater affinity for molec-
`ular oxygen. An appreciable volume change would be associ-
`ated with silicon removal from MoSi,, and it is suspected that
`
`GE-1022.004
`
`

`

`1944
`
`3000
`
`h % 2000
`8
`v .-
`F z
`g 1000
`
`U
`
`Journal of the American Ceramic SocietpSundaram et al.
`1500
`
`Vol. 78, No. 7
`
`1200
`
`h % 2 900
`Y
`.-
`F $ 600
`
`8-
`U
`
`300
`
`0
`0
`
`20
`
`40
`60
`Distance (pm)
`
`80
`
`100
`
`0
`0
`
`20
`
`40
`60
`Distance (pm)
`
`80
`
`100
`
`Fig. 6. Line EDS of molybdenum disilicide-glass interfacial region at (a) glass-line, (b) half-down regions.
`
`when a critical thickness has been reduced to Mo,Si,, it dis-
`lodges from the interface and floats away as debris. No longer
`near regions of MoSi,, local dissolved oxygen converts the
`Mo,Si, to Mo,Si, and Mo,Si to Mo, which was observed by
`XRD in regions in the glass extended away from the interface.
`The rate of corrosion was undoubtedly determined by the
`combined effects of the rates of coarsening of interfacial layers
`and the critical thicknesses at which interfacial stresses caused
`the coatings to crack and flake away, in turn, forcing new
`interfaces to form. Silicon diffusion through the Mo&
`is
`required to form vitreous silica, which dissolves into the glass.
`An appreciable volume percentage of silicon vacancies are
`expected to exist at the Mo,Si, interface; as silicon removal
`from the lattice continues, silicon vacancies form until a struc-
`tural collapse to a unit cell of different stoichiometry results.
`This is corroborated by the Si concentration gradiant shown by
`line EDS in Fig. 6. This type of interface would foster easy Si
`diffusion along these vacancy sites.
`
`(3) SI'C~,,~A~203
`Figures 8(a and b) show the SiC,,,/Al,O,-glass
`interfacial
`features at the glass-line and half-down regions, respectively.
`Bubbles formed within the glass, near the composite surface at
`both the glass-line and half-down regions. Near the glass line,
`the bubbles are seen in contact with the bulk of the specimen.
`CO or CO, can be the product of reaction between SIC and
`oxygen, depending on oxygen activity.', Competing oxidation
`reactions are considered in Fig. 9. The Sic portion of the
`composite oxidizes most favorably, to form silica and carbon
`monoxide. Since the CO-CO, equilibria fall below the Richard-
`son line (for Mo oxidation), the conversion of CO to CO, is not
`favorable, and the bubbles are expected to be purely carbon
`monoxide. Based on a standard Gibbs energy change of
`AGY,,,, = -534.67 kJ/mol, for the SiC,,,/CO,,, equilibria, and
`a CO bubble pressure of 1 atm, the molecular oxygen activity
`for the equilibria is uo, = 4.8 X
`Oxygen activities above
`this cause the reaction to be favorable.
`
`100
`
`0
`
`-100
`
`ka k
`-200
`a
`fi- W 0" -300
`
`a 3 -400
`3 g -500
`
`m 3 -600
`-700
`
`-800
`
`-900
`0
`
`.* 2 x 10-8
`
`1000
`2000
`Temperature (K)
`
`3000
`
`Fig. 7. Ellingham diagram of compounds of interest from MoSi,-oxygen interaction. A Richardson line of 2.0 X lo-' represents the highest
`possible molecular oxygen activity based on the results of Mo in E-glass. The dot-dashed line indicates the test temperature (1823 K). Thermody-
`namic data were taken from standard tab~lations.'~
`
`GE-1022.005
`
`

`

`July 1995
`
`Molten Glass Corrosion Resistance of Immersed Combustion-Heating Tube Materials in E-Glass
`
`1945
`
`SiC~,,/Al,O,-glass interfacial region after 48 h of corrosion at (a) glass-line and (b) half-down regions. C : composite, E: epoxy, and G: glass.
`Fig. 8.
`Left: secondary electron image. Right: backscattered image.
`
`The presence of bubbles in contact with a refractory in mol-
`ten glass has been demonstrated to be a mechanism of highly
`accelerated corrosion referred to as upward or downward
`“drilling,”’2 where convection driven by surface tension gradi-
`ents sweeps fresh glass to the interface and corrosion products
`away. Thus, oxygen-containing fresh glass would replace more
`viscous silica-rich glass. This corresponds well to the observed
`high corrosion rates for these specimens.
`(4) Remarks
`Oxidation did not occur with Mo immersed in E-glass, but
`did occur for MoSi, in the same glass. Mo,Si, debris converted
`to Mo,Si, which in turn converted to Mo. This could occur only
`for a soluble molecular oxygen activity of 2.4 X
`or higher.
`Thus, the oxygen activity of this composition of E-glass is
`and 2 X lo-’. This is an appre-
`bracketed between 2.4 X
`ciably lower oxygen activity than implied by corrosion studies
`in soda-lime-silicate glasses, where a soluble oxygen activity
`was based on the formation of a MOO, scale
`above 2.0 X
`on Mo.
`Immersed MoSi, corroded more slowly in E-glass than in
`soda-lime-silica glass, while the opposite was true for Sic,,,/
`
`A1,0,. It is interpreted that, with the lower oxygen activity in
`E-glass, the driving force for oxidation of MoSi, would be
`lessened. The 15°C higher temperature used in soda-lime-
`silicate glass studies may have also been a factor. The faster
`corrosion of SiC~,,/Al,O, in E-glass is attributed to the lower
`In a Iower
`viscosity of E-glass in this temperature
`viscosity glass, bubbles can more easily nucleate at the glass-
`solid interface and rise away from the interface. Removal of
`gaseous bubbles from the interface would keep product gases
`from suppressing the oxidation reaction.
`Extrapolation of these corrosion results to predicting the
`lifetime of components made of these materials in commercial
`glass tanks is mitigated by conditions unlike those of the experi-
`ment. So long as immersed Mo is not exposed to an E-glass
`which has random composition variances which would in turn
`permit increased molecular oxygen solubility, it should be
`indefinitely stable. For MoSi,, the convective flow forces com-
`mon in a glass tank may act to accelerate removal of the Mo,Si,
`interfacial layer and accelerate its corrosion behavior as com-
`pared to that presently reported. The atmosphere over the glass
`under combustion heating would be predominantly nitrogen,
`
`GE-1022.006
`
`

`

`1946
`
`Journal of the American Cei
`vamic Societ-undaram
`
`et al.
`
`Vol. 78, No. 7
`
`cracked away, fostering the formation of a new interfacial layer
`and continued specimen recession. SiC(,,/Al,O, receded at a
`rate approximately one order of magnitude greater, where the
`formation of CO bubbles fostered a mechanism of enhanced
`corrosion.
`
`References
`‘W. Lin, J. Y. Hsu, S. K. Sundaram, and R. F. Speyer, “Reactions between
`Flue Gas and Ceramic Composite Tubes for Radiant Heating”; presented at
`the 93rd Annual Meeting and Exposition of the American Ceramic Society,
`Cincinnati, OH, April 28-May 2, 1991 (Abstract No. 19-SII-91).
`*J. Y. Hsu, S. K. Sundaram, W. Lin, and R. F. Speyer, “Phase Equilibria
`Simulation between Immersed Radiant Tube Compositions and Molten Glass”;
`presented at the 93rd Annual Meeting and Exposition of the American Ceramic
`Society, Cincinnati, OH, April 28-May 2, 1991 (Abstract No. 38-SII-91).
`’S. K. Sundaram, J. Y. Hsu, and R. F. Speyer, “Molten Glass Corrosion
`Resistance of Immersed Combustion-Heating Tube Materials in Soda-Lime
`Silicate Glass,”J. Am. Ceram. Soc., 77 [6] 1613-23 (1994).
`*ASTM C621-84, “Standard Test Method for Isothermal Corrosion Resistance
`of Refractories to Molten Glass.” American Society for Testing and Materials,
`Philadelphia, PA, 1992.
`’L. Penberthy, “Electric Melting of Glass”; pp. 114-19 in Electric Melfing in
`the Glass Industry. Compiled by A. G. Pincus and G. M. Diken. Books for
`Industry and The Glass Industry Magazine, New York, 1976.
`‘R. Eck, “The State-of-the-Art of Molybdenum Fabrication,” Ceram. Eng.
`Sci. Proc., 6 [3-4] 274-86 (1985).
`’V. Bizzam, B. Linder, and N. Lindskog, “Molybdenum Disilicide Heating
`Elements: Meeting Advanced Ceramics Requirements,” Am. Cerum. Soc. Bull.,
`68 [lo] 1834-35 (1989).
`*S. K. Sundaram and R. F, Speyer, “Evaluation of Molten Glass Corrosion for
`Highly Corrosion Resistance Materials,” J. Test. Evul., in press.
`9Product Literature on Unicor501 AZS Refractory; p. Al. Corhart Refracto-
`ries, Louisville, KY, 1991.
`‘“R. H. Doremus, Glass Science; pp. 121-24. Wiley, New York, 1973.
`“K. Ooka, “Corrosion of Molybdenum by Molten Glasses,” J. Cerum. Assoc.
`Jpn., 72 [71 110-15 (1964).
`12V. L. Burdick, “The Corrosive Nature of Molten Glass”; pp. 545-61 in
`Introduction to Glass Science. Edited by L. D. Pye, H. J. Stevens, and W. C.
`Lacourse. Plenum Press, New York, 1972.
`‘’K. E. Spear, R. E. Tressler, Z. Zheng, and H. Du, “Oxidation of Silicon
`Carbide Single Crystals and CVD Silicon Nitride,” .I. Am. Cerum. Soc., 69 [9]
`674-81 (1986).
`’‘I. Barin, Thermochemical Data of Pure Substances, Part I and II; pp. 139,
`140, 157, 158, 292, 293, 301, 309, 929, 931-33, 1347-50, 1433-35, 1534-37,
`1687,1688, 1690. VCH, Weinheim, Germany, 1989.
`”E. J. Homyak, Owens-Illinois; private communication, 1990.
`I‘D. Miller, Owens-Coming Fiberglas; private communication, 1990.
`
`-700 ’
`
`0
`
`I
`I
`lo00
`2000
`Temperature (K)
`
`I
`3000
`
`Fig. 9. Ellingham diagram used to justify the chemistry of bubbles
`formed in the glass with oxidation of SiC,,,/Al,O,.
`
`water vapor, and carbon dioxide. This atmosphere would form
`less-oxidizing soluble species in the glass and may therefore
`allow MoSi, and SiC,,,/Al,O, to be stable in E-glass. With
`these caveats, the recession rates of MoSi, and SiC,,,/Al,O,
`components in E-glass, based on 24 h of corrosion, are 1.7 and
`38.6 cm/year, respectively.
`
`IV. Concf usion
`Mo showed no oxide layer formation when immersed in
`E-glass at 1550°C in an air atmosphere. MoSi, and SiC~,,/Al,O,
`showed reaction products indicating reactions with soluble
`molecular oxygen. A Mo,Si, layer formed which periodically
`
`GE-1022.007
`
`

This document is available on Docket Alarm but you must sign up to view it.


Or .

Accessing this document will incur an additional charge of $.

After purchase, you can access this document again without charge.

Accept $ Charge
throbber

Still Working On It

This document is taking longer than usual to download. This can happen if we need to contact the court directly to obtain the document and their servers are running slowly.

Give it another minute or two to complete, and then try the refresh button.

throbber

A few More Minutes ... Still Working

It can take up to 5 minutes for us to download a document if the court servers are running slowly.

Thank you for your continued patience.

This document could not be displayed.

We could not find this document within its docket. Please go back to the docket page and check the link. If that does not work, go back to the docket and refresh it to pull the newest information.

Your account does not support viewing this document.

You need a Paid Account to view this document. Click here to change your account type.

Your account does not support viewing this document.

Set your membership status to view this document.

With a Docket Alarm membership, you'll get a whole lot more, including:

  • Up-to-date information for this case.
  • Email alerts whenever there is an update.
  • Full text search for other cases.
  • Get email alerts whenever a new case matches your search.

Become a Member

One Moment Please

The filing “” is large (MB) and is being downloaded.

Please refresh this page in a few minutes to see if the filing has been downloaded. The filing will also be emailed to you when the download completes.

Your document is on its way!

If you do not receive the document in five minutes, contact support at support@docketalarm.com.

Sealed Document

We are unable to display this document, it may be under a court ordered seal.

If you have proper credentials to access the file, you may proceed directly to the court's system using your government issued username and password.


Access Government Site

We are redirecting you
to a mobile optimized page.





Document Unreadable or Corrupt

Refresh this Document
Go to the Docket

We are unable to display this document.

Refresh this Document
Go to the Docket